arxiv: v2 [astro-ph.co] 25 Jun 2012

Similar documents
The Los Cabos Lectures

Dwarf Galaxies as Cosmological Probes

Structure and substructure in dark matter halos

Princeton December 2009 The fine-scale structure of dark matter halos

Structure of Dark Matter Halos

arxiv: v1 [astro-ph] 5 Sep 2008

The mass of a halo. M. White

The subhalo populations of ΛCDM dark haloes

2. What are the largest objects that could have formed so far? 3. How do the cosmological parameters influence structure formation?

Structure formation in the concordance cosmology

arxiv: v2 [astro-ph.co] 18 Nov 2009

The Current Status of Too Big To Fail problem! based on Warm Dark Matter cosmology

The fine-scale structure of dark matter halos

Dark Matter Detection Using Pulsar Timing

The Formation and Evolution of Galaxy Clusters

The statistics of ΛCDM halo concentrations

Dependence of the inner dark matter profile on the halo mass

The Los Cabos Lectures

arxiv:astro-ph/ v1 27 Nov 2000

4. Structure of Dark Matter halos. Hence the halo mass, virial radius, and virial velocity are related by

4. Structure of Dark Matter halos. Hence the halo mass, virial radius, and virial velocity are related by

Some useful spherically symmetric profiles used to model galaxies

Gamma-rays from Earth-Size dark-matter halos

Simulating non-linear structure formation in dark energy cosmologies

The Caustic Technique An overview

THE BOLSHOI COSMOLOGICAL SIMULATIONS AND THEIR IMPLICATIONS

SUPPLEMENTARY INFORMATION

3 The lives of galaxies

The dark matter crisis

arxiv:astro-ph/ v3 13 May 2008

Cosmological Puzzles: Dwarf Galaxies and the Local Group

The Flatness of Mass-to-Light Ratio on Large Scales

Current status of the ΛCDM structure formation model. Simon White Max Planck Institut für Astrophysik

arxiv:astro-ph/ v1 12 Jul 1999

arxiv:astro-ph/ v2 19 Dec 2005

Testing the cold dark matter model with gravitational lensing

Massimo Meneghetti 1, Elena Torri 1, Matthias Bartelmann 2, Lauro Moscardini 3, Elena Rasia 1 and Giuseppe Tormen 1,

arxiv: v1 [astro-ph.co] 5 May 2017

The shape of cluster-scale dark matter halos

A Universal Density Profile from Hierarchical Clustering

arxiv: v1 [astro-ph.co] 26 Mar 2014

halo merger histories

arxiv:astro-ph/ v1 19 Nov 1999

You may not start to read the questions printed on the subsequent pages until instructed to do so by the Invigilator.

arxiv: v1 [astro-ph.co] 6 Sep 2012

Constraints on dark matter annihilation cross section with the Fornax cluster

Beyond the spherical dust collapse model

AST1100 Lecture Notes

Clusters of galaxies

IS THE NUMBER OF GIANT ARCS IN CDM CONSISTENT WITH OBSERVATIONS?

arxiv: v1 [astro-ph.co] 1 Oct 2009

A Gravitational Ising Model for the Statistical Bias of Galaxies

A5682: Introduction to Cosmology Course Notes. 11. CMB Anisotropy

arxiv: v1 [astro-ph] 31 Jul 2007

Galaxies 626. Lecture 3: From the CMBR to the first star

IoP. An Introduction to the Science of Cosmology. Derek Raine. Ted Thomas. Series in Astronomy and Astrophysics

Ensenada March 2008 Galaxy halos at (very) high resolution

The edge of darkness, and other halo surprises

Clusters: Observations

The Local Group Timing Argument

Cosmologists dedicate a great deal of effort to determine the density of matter in the universe. Type Ia supernovae observations are consistent with

FURTHER COSMOLOGY Book page T H E M A K E U P O F T H E U N I V E R S E

arxiv:astro-ph/ v4 18 Apr 2000

Astronomy 422. Lecture 15: Expansion and Large Scale Structure of the Universe

5.1 Circular Velocities and Rotation Curves

The spatial distribution of galactic satellites in the cold dark matter cosmology

Milky Way s Mass and Stellar Halo Velocity Dispersion Profiles

arxiv:astro-ph/ v1 24 Feb 2005

The Millennium Simulation: cosmic evolution in a supercomputer. Simon White Max Planck Institute for Astrophysics

pseudo- evolution of halo mass and observable-mass relations Andrey Kravtsov The University of Chicago

Signal Model vs. Observed γ-ray Sky

Galaxy interaction and transformation

N-body Simulations. Initial conditions: What kind of Dark Matter? How much Dark Matter? Initial density fluctuations P(k) GRAVITY

Observational Cosmology

Infrared Mass-to-Light Profile Throughout the Infall Region of the Coma Cluster

Cosmological Simulations: Successes & Tensions of ɅCDM

(Toward) A Solution to the Hydrostatic Mass Bias Problem in Galaxy Clusters. Eiichiro Komatsu (MPA) UTAP Seminar, December 22, 2014

Galaxy clusters. Dept. of Physics of Complex Systems April 6, 2018

Components of Galaxies: Dark Matter

arxiv:astro-ph/ v1 10 Nov 1999

The halo-galaxy connection KITP, May 2017` Assembly bias. Simon White Max Planck Institute for Astrophysics

How many arcmin-separation lenses are expected in the 2dF QSO survey?

A brief history of cosmological ideas

The motion of emptiness

Dark matter from cosmological probes

arxiv: v1 [astro-ph.co] 12 Oct 2010

Chapter 9. Cosmic Structures. 9.1 Quantifying structures Introduction

X-ray and Sunyaev-Zel dovich Effect cluster scaling relations: numerical simulations vs. observations

Gaia Revue des Exigences préliminaires 1

ASTRON 449: Stellar (Galactic) Dynamics. Fall 2014

AS1001:Extra-Galactic Astronomy

Dark Energy vs. Dark Matter: Towards a unifying scalar field?

arxiv: v1 [gr-qc] 1 Dec 2017

In Quest of a True Model of the Universe 1

Surface Brightness of Spiral Galaxies

Dark Matter Density Function Depending on Gravitational Field as Universal Law - M. Abarca

OBSERVATIONAL EVIDENCE FOR DARK MATTER AND DARK ENERGY. Marco Roncadelli INFN Pavia (Italy)

Newtonian Gravity and Cosmology

Additional on-line material for paper A comprehensive set of simulations of high-velocity collisions between main sequence stars.

Clusters: Observations

Transcription:

Mon. Not. R. Astron. Soc. 000, 000 000 (0000) Printed 17 February 2013 (MN LATEX style file v2.2) The Phoenix Project: the Dark Side of Rich Galaxy Clusters arxiv:1201.1940v2 [astro-ph.co] 25 Jun 2012 L. Gao 1,2, J. F. Navarro 3, C. S. Frenk 2, A. Jenkins 2, V. Springel 4,5, S. D. M. White 6 1 Partner Group of the Max Planck Institute for Astrophysics, National Astronomical Observatories, Chinese Academy of Sciences, Beijing, 100012, China 2 Institute of Computational Cosmology, Department of Physics, University of Durham,Science Laboratories, South Road, Durham DH1 3LE 3 Department of Physics and Astronomy, University of Victoria, PO Box 3055 STN CSC, Victoria, BC, V8W 3P6 Canada 4 Heidelberger Institut für Theoretische Studien, Schloss-Wolfsbrunnenweg 35, 69118 Heidelberg, Germany 5 Zentrum für Astronomie der Universität Heidelberg, Astronomisches Recheninstitut, Mönchhofstr. 12-14, 69120 Heidelberg, Germany 6 Max-Planck Institute for Astrophysics, Karl-Schwarzschild Str. 1, D-85748, Garching, Germany 17 February 2013 ABSTRACT We introduce the Phoenix Project, a set of ΛCDM simulations of the dark matter component of nine rich galaxy clusters. Each cluster is simulated at least at two different numerical resolutions. For eight of them, the highest resolution corresponds to 130 million particles within the virial radius, while for one this number is over one billion. We study the structure and substructure of these systems and contrast them with six galaxy-sized dark matter haloes from the Aquarius Project, simulated at comparable resolution. This comparison highlights the approximate mass invariance of CDM halo structure and substructure. We find little difference in the spherically-averaged mass, pseudo-phase-space density, and velocity anisotropy profiles of Aquarius and Phoenix haloes. When scaled to the virial properties of the host halo, the abundance and radial distribution of subhaloes are also very similar, despite the fact that Aquarius and Phoenix haloes differ by roughly three decades in virial mass. The most notable difference is that cluster haloes have been assembled more recently and are thus significantly less relaxed than galaxy haloes, which leads to decreased regularity, increased halo-to-halo scatter and sizable deviations from the mean trends. This accentuates the effects of the strong asphericity of individual clusters on surface density profiles, which may vary by up to a factor of three at a given radius, depending on projection. The high apparent concentration reported for some strong-lensing clusters might very well reflect these effects. A more recent assembly also explains why substructure in some Phoenix haloes is slightly more abundant than in Aquarius, especially in the inner regions. Resolved subhaloes nevertheless contribute only 11 ± 3% of the virial mass in Phoenix clusters. Together, the Phoenix and Aquarius simulation series provide a detailed and comprehensive prediction of the cold dark matter distribution in galaxies and clusters when the effects of baryons can be neglected. Key words: methods: N-body simulations methods: numerical dark matter galaxies: haloes galaxies:clustering 1 INTRODUCTION The past two decades have witnessed the emergence of a paradigm for the origin of structure in the Universe. There is now strong evidence that the dominant forms of the matter-energy content are a combination of a mysterious form of dark energy that governs the late expansion of the Universe, and dark matter made up of some kind of non-baryonic, weakly interacting elementary particle left over from the Big Bang. Although the exact nature of the dark matter particle is unknown, astrophysical clues to its identity may be gained by studying its clustering properties on different scales. Considerable effort has been devoted to this task, and has led to the Email:lgao@bao.ac.cn crafting of detailed theoretical predictions, especially for the case of particles with negligible thermal velocity, the cornerstone of the popular cold dark matter (CDM) theory. As a result, we now understand fairly well: (i) the statistics of CDM clustering on large scales and its dependence on the cosmological parameters (e.g., Jenkins et al. 1998; Springel et al. 2006); (ii) the dynamics of its incorporation into non-linear units ( haloes ) (see, e.g., Wang et al. 2011, and references therein); and, at least empirically, (iii) its spatial distribution within such virialized structures (e.g., Frenk et al. 1985; Navarro et al. 1996, 1997, 2004, 2010). Progress in this field has been guided by N-body simulations of ever increasing numerical resolution and dynamic range (e.g. Frenk et al. 1985; Navarro et al. 1997; Moore et al. 1999; Jing & Suto 2002; Navarro et al. 2004; Gao et al. 2004b;

2 Gao et al. Diemand et al. 2004a, 2007; Gao et al. 2008; Springel et al. 2008a; Stadel et al. 2009; Navarro et al. 2010). These simulations are essential to investigate highly non-linear scales such as the haloes of individual galaxies and galaxy groupings, where simple analytical approximations fail. A few key properties of CDM haloes are now widely agreed upon, at least when the effects of baryons are neglected: (a) the presence of a central density cusp ; (b) strong deviations from spherical symmetry; (c) a remarkable similarity in the shape of the mass profiles; and (d) the presence of abundant substructure in the form of self-bound subhaloes. On the scale of individual galaxies, these key predictions have been confirmed and refined by the latest simulation work, in particular the Via Lactea simulation series (Diemand et al. 2007), the GHALO simulation (Stadel et al. 2009) and the Aquarius Project (Springel et al. 2008b,a; Navarro et al. 2010). For example, the central density cusp is now accepted to be shallower than hypothesized in some earlier work and mass profiles have been shown to be only approximately self-similar. Further, it is now clear that although subhaloes are subdominant in terms of total mass, they are still dense and abundant enough to dominate the dark matter annihilation radiation from a halo. As shown by Springel et al. (2008a), the latter statement requires a detailed characterization of the substructure, including the internal properties of the subhaloes, their mass function, and their spatial distribution within the main halo. The Aquarius Project has provided compelling, if mainly empirical, guidance to each of these issues in the case of haloes similar to that of the Milky Way. For example, the subhalo mass function is well approximated by a power law, dn/dm M 1.9, with a normalisation, in scaled units, weakly dependent on halo mass (Gao et al. 2011a). In addition, subhaloes tend to avoid the central region of the main halo and are more prevalent in the outer regions. Interestingly, their spatial distribution appears independent of subhalo mass, a result that, if generally applicable, simplifies substantially the characterization of substructure. Finally, the internal structure of subhaloes obeys scaling laws similar to those of haloes in isolation but slightly modified by the effects of the tidal field of the main halo: subhaloes are denser, reaching their peak circular velocity at radii roughly half that of their isolated counterparts. Galaxy clusters are a promising venue for testing these predictions. The central cusp, for example, can be constrained by combining measurements of the stellar kinematics of the central galaxy with a lensing analysis of radial and tangential arcs near the cluster centre (e.g., Sand et al. 2002, 2004; Meneghetti et al. 2007; Newman et al. 2009; Zitrin et al. 2011). Outside the very centre, the cluster mass profile can be measured through weak lensing (see, e.g., Okabe et al. 2010; Oguri et al. 2011; Umetsu et al. 2011), X-ray studies of the hot intracluster medium (ICM; e.g, Buote et al. 2007), and, more recently, through the ICM Sunyaev- Ze ldovich effect on the cosmic microwave background (see, e.g., Gralla et al. 2011). In many cases, including substructure seems required in order to obtain acceptable fits (e.g. Mao & Schneider 1998; Mao et al. 2004; Xu et al. 2009; Natarajan et al. 2007, 2009), implying that it should be possible to contrast observations directly with the CDM substructure predicted by simulations. Such endeavour has so far been hindered by the lack of ultrahigh-resolution dark matter simulations of galaxy clusters comparable to the Aquarius series. Indeed, the highest-resolution galaxy cluster simulations published to date have at most of order a few million particles within the virial radius (e.g. Jing & Suto 2000; Springel et al. 2001a; Diemand et al. 2004a; Reed et al. 2005), roughly one thousand times fewer than the best resolved Aquarius halo. None of these cluster simulations are thus able to address conclusively issues such as the structure of the central cusp or the properties of cluster substructure. Although it may be tempting to appeal to the nearly selfsimilar nature of CDM haloes to extrapolate the Aquarius results to cluster scales, it is unclear what systematic uncertainties might be introduced through such extrapolation. Clusters are rare, dynamically young objects up to one thousand times more massive than individual galaxies. They thus trace scales where the CDM power spectrum differs qualitatively from that of galaxies. Precision work demands that the near self-similarity of dark haloes be scrutinized directly in order to provide definitive predictions for the CDM paradigm on cluster scales. To this aim, we have carried out a suite of simulations designed to address these issues in detail. The Phoenix Project follows the design of the Aquarius Project and consists of zoomed-in resimulations of individual galaxy clusters drawn from a cosmologically representative volume. The simulations follow only the dark matter component of each cluster, and include the first simulation of a cluster-sized halo with more than one billion particles within the virial radius. Like the Aquarius Project on galaxy scales, the large dynamic range of these simulations allows us to probe not only the innermost regions of cluster haloes and thus the structure of the central cusp, but also the statistics, internal structure, and spatial distribution of cluster substructure over a mass range spanning seven decades. Our paper is organized as follows. In Section 2, we describe our numerical techniques and introduce the simulation set. In Sec. 3 and Sec. 4 we discuss, respectively, the density profile and substructure properties of Phoenix haloes and compare them with those of Aquarius. Sec. 5 summarizes our main conclusions. 2 THE SIMULATIONS The Phoenix Project consists of a series of simulations of 9 different galaxy clusters with masses exceeding 5 10 14 h 1 M. These clusters were selected from a large cosmological volume and resimulated individually at varying resolution. Details of the resimulation procedure are given below. 2.1 Cosmology All the simulations reported here adopt the cosmological parameters of the Millennium Simulation (Springel et al. 2005): Ω M = 0.25, Ω Λ = 0.75, σ 8 = 0.9, n s = 1, and a present-day value of the Hubble constant H 0 = 100hkm s 1 Mpc 1 = 73 km s 1 Mpc 1. This is also the set of cosmological parameters adopted for the Aquarius project (Springel et al. 2008b), which targeted haloes a thousand times less massive. Although they are inconsistent with the latest CMB data (Komatsu et al. 2011) the differences are not large (the main difference is that a lower value of σ 8 = 0.81 is now preferred) and they are expected to affect only the abundance of cluster haloes rather than their detailed structure and substructure properties (Wang et al. 2012). This choice also has the advantage that any difference between Aquarius and Phoenix haloes can be traced to the different mass scales and not to variations in the cosmological model.

The Phoenix project 3 Name m p M 200 r 200 N 200 ε r conv [h 1 M ] [h 1 M ] [h 1 Mpc] [h 1 kpc] [h 1 kpc] Ph-A-1 6.355 10 5 6.560 10 14 1.413 1,032,269,120 0.15 1.2 5.084 10 6 6.570 10 14 1.414 129,235,472 0.32 2.7 Ph-A-3 1.716 10 7 6.566 10 14 1.413 38,261,560 0.7 4.2 Ph-A-4 1.373 10 8 6.593 10 14 1.415 4,802,516 2.8 9.4 Ph-B-2 6.127 10 6 8.255 10 14 1.526 134,718,112 0.32 3.0 Ph-B-4 1.656 10 8 8.209 10 14 1.522 4,956,688 2.8 10.7 Ph-C-2 4.605 10 6 5.495 10 14 1.386 119,324,008 0.32 2.6 Ph-C-4 1.182 10 8 5.549 10 14 1.383 4,696,046 2.8 9.2 Ph-D-2 4.721 10 6 6.191 10 14 1.386 130,529,200 0.32 2.7 Ph-D-4 1.373 10 8 6.162 10 14 1.384 4,488,330 2.8 9.4 Ph-E-2 4.425 10 6 5.969 10 14 1.369 130,529,200 0.32 2.4 Ph-E-4 1.017 10 8 5.923 10 14 1.366 5,824,375 2.8 8.4 Ph-F-2 4.425 10 6 7.997 10 14 1.509 129,221,216 0.32 2.8 Ph-F-4 1.682 10 8 8.039 10 14 1.512 4,779,008 2.8 10.3 Ph-G-2 8.599 10 6 1.150 10 15 1.704 133,730,958 0.32 3.2 Ph-G-4 2.907 10 8 1.148 10 15 1.703 3,949,310 2.8 13.1 Ph-H-2 8.600 10 6 1.136 10 15 1.686 129,488,456 0.32 2.9 Ph-H-4 2.502 10 8 1.150 10 15 1.686 4,456,720 2.8 11.8 Ph-I-2 1.841 10 7 2.411 10 15 2.185 131,845,620 0.32 2.9 Ph-I-4 4.559 10 8 2.427 10 15 2.181 5,289,259 2.8 14.2 Table 1. Basic parameters of the Phoenix simulations. Each of the nine haloes is labelled as Ph-X-N, where the letter X (from A to I) identifies each halo, and N, which runs from 1 to 4, refers to the numerical resolution (1 is highest). The parameter m p gives the particle mass in the high-resolution region that includes the cluster; M 200 is the virial mass of the halo; r 200 is the corresponding virial radius; and N 200 states the number of particles inside r 200. The parameter ε is the Plummer-equivalent gravitational softening length, so that pairwise interactions are fully Newtonian when separated by a distance greater than 2.8ε. The last column lists the convergence radius, r conv, outside which the circular velocity is expected to converge to better than 10%. 2.2 Cluster Sample and Resimulations The Phoenix cluster sample is selected for resimulation from the Millennium Simulation friends-of-friends group catalog at z = 0. Six clusters were selected at random from the 72 systems with virial 1 mass in the range 5 < M 200 /10 14 h 1 M < 10. In order to sample the tail of rare rich clusters three further Phoenix clusters were selected from the nine Millennium halos which have M 200 > 10 15 h 1 M. The initial conditions for resimulation were set up using a procedure analogous to that used for the Aquarius haloes and described in detail by Power et al. (2003) and Springel et al. (2008a). The only difference is that the initial displacements and velocities were computed using second-order Lagrangian perturbation theory, as described by Jenkins (2010). All nine haloes were resimulated at least twice using different numerical resolution (level 2 and level 4, respectively). At level 2 each cluster has between 120 and 135 million particles within the virial radius; at level 4 each system is made up of 4 to 6 million particles. We have selected one of the clusters (Ph-A) for a numerical resolution study and have carried out an extra level-3 run (with roughly 40 million particles within r 200 ) and a flagship level-1 run, where we followed 4.05 billion high-resolution particles in 1 We define the virial radius of a cluster, r 200, as that of a sphere of mean density 200 times the critical density for closure; ρ crit = 3H0 2 /8πG. The virial radius defines implicitly the virial mass of a cluster, M 200, and its virial velocity, V 200 = GM 200 /r 200. total, 1.03 billion of which are found within r 200 at z = 0. For ease of reference we label the runs using the convention Ph-X- N, where X is a letter from A to I that identifies each individual cluster and N is a number from 1 to 4 that specifies the resolution level. The simulation parameters are summarized in Table 1. We have used for all runs the P-Gadget-3 code, a version of Gadget- 2 (Springel et al. 2005) especially optimized for zoomed-in cosmological resimulations in distributed-memory massively-parallel computers. The code is identical to that used for the Aquarius Project (Springel et al. 2008b). The simulations were carried out on Deepcomp 7000 at the Supercomputer Center of the Chinese Academy of Science. The largest simulation, Ph-A-1, used 3 Tbs of memory on 1024 cores and took about 1.9 million CPU hours. The initial conditions were generated at the Institute for Computational Cosmology (Durham University). The gravitational softening of each run was chosen following the optimal prescription of Power et al. (2003). It is kept fixed in comoving coordinates throughout each run and is listed in Table 1. Our highest-resolution run (Ph-A-1) has a nominal (Plummerequivalent) spatial resolution of just 150h 1 pc. Haloes are identified in each run using the friends-of-friends (FOF) group finding algorithm with linking length set to 20% of the mean interparticle separation (Davis et al. 1985). Substructure within FOF haloes is identified by SUBFIND (Springel et al. 2001b), a groupfinding algorithm that searches recursively for self-bound subhaloes. Both FOF and SUBFIND have been integrated within P- Gadget-3 and are run on-the-fly each time a simulation snapshot is created.

4 Gao et al. Figure 1. Images of cluster Ph-A at four different numerical resolutions. Each panel projects a cubic volume 5h 1 Mpc on a side. The brightness of each image pixel is proportional to the logarithm of the square of the dark matter density projected along the line of sight, and the hue encodes the local velocity dispersion density-weighted along the line of sight (see text for details). This rendering choice highlights the presence of substructure which, although abundant, contributes less than about 10% of the total mass within the virial radius. We have stored for each run 72 snapshots uniformly spaced in log 10 a, starting at a= 0.017 (a=1/(1+z) is the expansion factor). The initial conditions are set at z init = 63 for our level 4 and at z init = 79 for the rest. The large number of outputs is designed to allow us in future work to implement semi-analytic models of galaxy formation in order to follow the evolution of the baryonic component of galaxies within rich clusters. We list the basic structural parameters of Phoenix clusters at redshift z = 0 in Table 2. These include the peak circular velocity, V max, and the radius, r max, at which it is reached; the half-mass formation redshift, z h, when the main progenitor first reaches half the final halo mass; the concentration parameters, c and c E, obtained from the best-fit NFW (Navarro et al. 1996, 1997) and Einasto (1965) profiles, respectively; the figure of merit, Q NFW and Q E, associated with each of those fits; and the Einasto shape parameter α. (See the Appendix for definitions corresponding to these fitting formulae and for details on the profile-fitting procedure.) N sub is the total number of subhaloes with more than 20 particles identified by SUBFIND inside r 200 ; f sub is the total mass contributed by these subhaloes, expressed as a fraction of the virial mass. 3 THE STRUCTURE OF PHOENIX CLUSTERS We shall focus our analysis on the properties of Phoenix clusters at z = 0. Figure 1 shows Ph-A at the four different numerical resolutions. As in Springel et al. (2008a), this and other cluster images are constructed so that the brightness of each pixel is proportional

The Phoenix project 5 Figure 2. The inner 1h 1 Mpc of Ph-A-1. Color coding is as in Fig. 1. This figure illustrates clearly the strong asphericity of the halo; the presence of several nested levels of substructure, and the tendency of subhaloes to avoid the halo centre. to the logarithm of the square of the dark matter density projected along the line of sight, S(x,y)= ρ 2 loc (r)dz (1) while the color hue encodes the mean dark matter velocity dispersion, σ(x,y)= 1 S(x, y) σ loc (r)ρ 2 loc (r)dz (2) Here the local dark matter density, ρ loc (r), and the local velocity dispersion, σ loc (r), are estimated using an SPH kernel interpolation scheme. Figure 1 shows that the main result of increasing the number of particles is the ability to resolve larger numbers of subhaloes. On the other hand, the main properties of the cluster, such as its shape and orientation, the overall mass profile, and even the location of the largest subclumps remain invariant in all four Ph-A realizations. Fig. 2 is analogous to Fig. 1, but for the inner 1h 1 Mpc of Ph-A-1 (our highest resolution run). This image highlights the strong asphericity of the halo, as well as the presence of several nested levels of substructure (i.e., subhaloes within subhaloes). It also shows that subhaloes tend to avoid the central regions. These characteristics are shared with galaxy-sized haloes (Springel et al. 2008a), and appear to be typical of CDM haloes on all mass scales. Fig. 3 is analogous to Fig. 1 but for all level-2 Phoenix haloes at z=0. This figure shows that the main characteristics of Ph-A described above are common to all Phoenix clusters: strong asphericity; abundant substructure; and a marked difference between the spatial distribution of mass (which is highly concentrated) and that of subhaloes (which tend to avoid the central regions).

6 Gao et al. Figure 3. As Fig. 1, but for all level-2 Phoenix clusters at z=0. Boxes are all 5h 1 Mpc on a side. Note that the appearance of several Phoenix clusters is suggestive of a transient evolutionary stage, characterized by the presence of a number of undissolved substructure groupings. Ph-G-2 is a particularly good example of this irregular structure which may be traced to its recent assembly time; this cluster has acquired half its mass since z = 0.18. Fig. 3 also highlights an important characteristic of clustersized dark matter haloes: the presence of multiple centres traced by groups of subhaloes, as well as the overall impression that many systems are in a transient, unrelaxed stage of their evolution. This is expected, given the late assembly of the clusters: Ph-G-2, for example, assembled half its final mass after z = 0.18; the median half-mass assembly redshift for all Phoenix clusters is just z = 0.56. Ph-A, on the other hand, appears relaxed; this cluster has the highest formation redshift of our sample, z h 1.2. The late assembly and concomitant departures from equilibrium are characteristics that set clusters apart from galaxy-sized haloes; for comparison, the median half-mass formation redshift of Aquarius haloes is z 2. Table 2 lists two quantitative measures of departures from equilibrium: the fraction of mass in substructures, f sub, and the offset, d off, between the centre of mass of the halo and the location of the potential minimum expressed in units of the virial radius (for further discussion of these parameters see Neto et al. 2007). These correlate well with the formation redshift, z h, and are significantly larger, on average, than in the galaxy-sized Aquarius haloes (see Table 3). 3.1 Mass Profiles We explore in this section the spherically-averaged mass profiles of Phoenix haloes. We begin by using the four Ph-A realizations to assess the limitations introduced by finite numerical resolution. Fig. 4 shows the density profile, ρ(r), as well as the radial dependence of the logarithmic slope, γ = d ln ρ/d ln r, for Ph-A-1 through Ph-A- 4. As discussed by Power et al. (2003) and Navarro et al. (2010), the mass profiles of simulated haloes are robustly determined in

The Phoenix project 7 1000.0 3.5 3.0 100.0 2.5 ρ r 2 [10 10 M O Mpc] 10.0 Ph-A-4 Ph-A-3 Ph-A-1 γ=-dlnρ/dlnr 2.0 1.5 1.0 Einasto fits NFW fits 1.0 0.1 0.0001 0.0010 0.0100 0.1000 1.0000 r [h -1 Mpc] 0.5 0.0001 0.0010 0.0100 0.1000 1.0000 r [h -1 Mpc] Figure 4. Left panel: Spherically-averaged density profile of halo Ph-A at z = 0. Different colors correspond to the four different resolution runs listed in Table 1. The panel on the left shows the density multiplied by r 2 in order to enhance the dynamic range of the plot. Each profile is shown with a thick line connecting filled circles from the convergence radius, r conv, outwards (Power et al. 2003). Thin curves extend the profiles inwards down to r = 2ε, where ε is the Plummer-equivalent gravitational softening length. Vertical dotted lines indicate, for each run, 2.8ε, the distance beyond which pairwise particle interactions are fully Newtonian. Note the excellent numerical convergence achieved for each simulation outside their r conv. An NFW profile with concentration c=5.63 (thin dashed brown line) and an Einasto profile with α=0.22 and c E = 5.59 (thin dashed magenta line) are also shown for comparison. Right panel: Logarithmic slope (γ = d lnρ/d ln r) of the density profile as a function of radius. Colors and line types are the same as in the left panel. Note again the excellent convergence achieved in all runs at radii outside the convergence radius, r conv. Name V max r max z h c E c Q E Q NFW α N sub f sub d off [km s 1 ] [h 1 Mpc] Ph-A-1 1521.82 0.55 1.17 5.59 5.63 0.037 0.093 0.215 192,206 0.080 0.04 1527.24 0.55 1.17 5.72 5.96 0.039 0.075 0.216 26,896 0.071 0.04 Ph-A-3 1529.41 0.56 1.17 5.69 6.04 0.038 0.061 0.218 8,478 0.062 0.04 Ph-A-4 1538.88 0.59 1.17 5.71 6.14 0.052 0.063 0.219 1,049 0.049 0.04 Ph-B-2 1624.52 0.53 0.46 4.41 4.19 0.127 0.108 0.235 38,659 0.108 0.02 Ph-B-4 1623.12 0.56 0.46 4.40 4.06 0.107 0.117 0.276 1,657 0.081 0.02 Ph-C-2 1294.19 0.65 0.76 4.27 5.11 0.077 0.104 0.181 33,529 0.114 0.06 Ph-C-4 1310.19 0.78 0.76 4.34 4.72 0.085 0.112 0.185 1,489 0.095 0.06 Ph-D-2 1393.13 0.68 0.46 3.88 4.08 0.122 0.086 0.205 38,199 0.124 0.05 Ph-D-4 1436.10 0.65 0.46 4.03 4.34 0.136 0.127 0.212 1,491 0.093 0.05 Ph-E-2 1385.78 0.65 0.91 3.48 5.19 0.067 0.135 0.149 33,678 0.101 0.04 Ph-E-4 1399.96 0.68 0.91 4.02 4.82 0.048 0.079 0.181 1,547 0.070 0.04 Ph-F-2 1543.27 0.60 1.1 3.81 4.61 0.053 0.048 0.186 31,247 0.095 0.05 Ph-F-4 1559.44 0.62 1.1 4.00 4.54 0.059 0.057 0.203 1,547.075 0.05 Ph-G-2 1561.75 1.06 0.18 0.78 3.33 0.100 0.221 0.097 42,528 0.168 0.17 Ph-G-4 1599.17 1.04 0.18 1.10 2.98 0.109 0.164 0.116 1,586 0.140 0.17 Ph-H-2 1676.43 1.14 0.21 1.98 4.66 0.155 0.212 0.117 35,048 0.095 0.1 Ph-H-4 1710.19 1.14 0.21 2.75 3.59 0.109 0.115 0.178 1,437 0.069 0.1 Ph-I-2 2236.05 1.03 0.56 4.18 4.86 0.041 0.059 0.190 35,754 0.102 0.02 Ph-I-4 2269.09 1.05 0.56 4.48 5.02 0.045 0.051 0.208 1,641 0.073 0.02 Table 2. Basic structural parameters of Phoenix clusters at z=0. The leftmost column labels each run, as in Table 1; the second and third columns list the peak circular velocity, V max, and the radius, r max, at which it is reached. The concentration parameters of the best NFW (Navarro et al. 1996, 1997) and Einasto (Einasto 1965) fits are listed under c and c E, respectively. Q NFW and Q E are the figures of merit of the best NFW and Einasto fits, respectively. The column labelled α lists the Einasto shape parameter. N sub denotes the total number of subhaloes with more than 20 particles identified within r 200 ; f sub is the fraction of the virial mass contributed by such subhaloes; and d off is the distance from the gravitational potential minimum to the centre of mass of particles within the virial radius, in units of r 200.

8 Gao et al. Name z h d off f sub α Q min C χ χ N m s N v d Phoenix 0.65 0.061 0.109 0.175 0.086 1.75 1.86 7866 0.97 3984 3.32 ±0.36 ±0.047 ±0.027 ±0.046 ±0.041 ±0.29 ±0.04 ±965 ±0.02 ±317 ±0.10 Aquarius 1.65 0.032 0.071 0.159 0.048 2.19 1.82 5092 0.94 4033 3.13 ±0.65 ±0.011 ±0.022 ±0.022 ±0.012 ±0.14 ±0.02 ±677 ±0.02 ±500 ±0.09 Table 3. Comparison of the average properties of the six galaxy-sized Aquarius haloes and the nine cluster-sized Phoenix haloes. Sample averages are listed for each quantity together with the rms dispersion around the mean. The first column identifies the simulation set; z h is the half-mass formation redshift; d off and f sub are the dynamical relaxation diagnostics introduced in Table 2; α is the best-fit Einasto shape parameter and Q min the goodness of fit measure (Sec. 6); C χ and χ are the parameters of power-law fits to the pseudo-phase-space density profile (eq. 3); N m and s describe the power-law fits to the subhalo mass function, N(> µ)=n m (µ/10 6 ) s (eq. 4); N v and d those corresponding to fits of the form, N(> ν)=n v (ν/0.025) d, to the subhalo velocity function (eq. 5). 3.5 ρ/ρ c (r/r 200 ) 2 100 10 γ=-dlnρ/dlnr 3.0 2.5 2.0 1.5 Ph-B-2 Ph-C-2 Ph-D-2 Ph-E-2 Ph-F-2 Ph-G-2 Ph-H-2 Ph-I-2 <Phoenix> <Aquarius> 1.0 1 0.001 0.010 0.100 1.000 r/r 200 0.5 0.001 0.010 0.100 1.000 r/r 200 3.5 1.00 3.0 2.5 ρ/ρ -2 r 2 2 /r -2 0.10 γ=dlnρ/dlnr 2.0 1.5 1.0 0.01 0.01 0.10 1.00 10.00 r/r -2 0.5 0.01 0.10 1.00 10.00 r/r -2 Figure 5. Spherically-averaged density (left panels) and logarithmic slope (right panels) of all level-2 Phoenix haloes as a function of radius. Radii have been scaled to the virial radius of each halo in the top panels and to the scale radius, r 2, of the best-fit Einasto profile in the bottom panels. Profiles are plotted down to the convergence radius, r conv. The thick dashed black line shows the average density profile of all Phoenix haloes, computed after stacking the nine haloes, each scaled to its own virial mass and radius. The thick red dashed line shows the result of the same stacking procedure, but applied to the Aquarius haloes. regions where the two body-relaxation time exceeds the age of the Universe. This constraint defines a convergence radius, r conv, outside which the circular velocity, V c = (GM(< r)/r) 1/2, is expected to converge to better than 10%. Since V c is a cumulative measure we expect r conv to be a conservative indicator of the innermost radius where local estimates of the density, ρ(r), converge to better than 10%. This is indeed the case for Ph-A, as shown in Fig. 4. The left panel shows ρ(r), multiplied by r 2 in order to remove the dominant radial trend so as to enhance the dynamic range of the plot. Thick lines highlight the radial range of the profile outside the convergence radius; the density clearly converges to better than 10% at radii greater than r conv. In those regions the logarithmic slope γ is also robustly and accurately determined. We conclude that r> r conv is a simple and useful prescription that identifies the regions unaffected by numerical limitations. We list r conv for all Phoenix runs in Table 1. The thin dashed lines in Fig.4 indicate the best-fit NFW

The Phoenix project 9 (brown) and Einasto (magenta) profiles, computed as described in the Appendix. The NFW shape is fixed in this log-log plot, whereas the Einasto shape is controlled by the parameter α, which is found to be 0.215 by the fitting procedure when applied to the Ph-A-1 profile. This figure suggests that the shape of the mass profile deviates slightly but systematically from the NFW profile. Although it is possible to obtain excellent fits over the resolved radial range with the NFW formula (typical residuals are less than 10%), there is a clear indication that the density profile near the centre is shallower than the asymptotic r 1 NFW cusp. In agreement with results from the Aquarius Project (Navarro et al. 2010), there is little indication that the central density cusp of Ph-A is approaching a power-law; the profile becomes gradually shallower all the way in to the innermost resolved radius. This radial trend is very well described by the Einasto profile. Fig. 5 shows the density profiles of all level-2 Phoenix haloes, in a format similar to that of Fig. 4. The top panels show profiles with radii scaled to the virial radius of each cluster, whereas those at the bottom show radii in units of the scale radius, r 2, of the best Einasto fit. Profiles are shown from the convergence radius, r conv, outwards. In general, the density profiles of Phoenix clusters become gradually shallower towards the centre; from γ 3in the outer regions to an average value of γ 1 at the innermost resolved radius. This behaviour is similar to that of Aquarius haloes, whose average profile is shown by the thick red dashed lines in Fig. 5. The large difference between Aquarius and Phoenix seen in the top panels of this figure just reflect the different concentration 2 of cluster- and galaxy-sized haloes. Indeed, when radii are scaled to r 2, the average Phoenix and Aquarius profiles are basically indistinguishable from each other. This is confirmed quantitatively by the best-fit Einasto parameters of these average profiles (listed in Table 3). The average Phoenix halo is only slightly worse fit by an Einasto profile than Aquarius, as shown by the Q min goodness-of-fit measure (6.5% vs 1.8%, respectively). There is also a slight difference in shape parameter; the average Phoenix cluster has α = 0.175 whereas the average Aquarius halo has α = 0.159, in agreement with previously reported trends (Gao et al. 2008). One aspect in which Phoenix and Aquarius haloes do differ is the halo-to-halo scatter: the dispersion in the Einasto parameter α is twice as large for clusters as for galaxy-sized haloes (Table 3). This may be readily seen in Fig. 5:, for example, follows the steady decline in γ towards the centre exhibited by Ph-A-1 (and characteristic also of Aquarius haloes), whereas in other cases, such as Ph-H-2, γ stays roughly constant over a wide radial range near the centre. The latter behaviour is poorly captured by the Einasto or NFW fitting formulae, and leads to larger residuals and figure-of-merit values for the best fits. NFW and Einasto best-fit residuals are shown in Fig. 6; per bin deviations of up to 40% from NFW and 20% from Einasto fits are not uncommon for Phoenix clusters. These deviations may be traced to transient departures from equilibrium induced by the recent formation of many Phoenix clusters. For example, one of the worst offenders is Ph-H-2, which accreted half its final mass since z = 0.21 and whose unrelaxed appearance is obvious in Fig. 3. In contrast,, the cluster with 2 The concentration is defined as r 200 /r 2, where r 2 is the radius at which the logarithmic slope γ has the isothermal value of 2. This indicates the location of the maximum of the curves shown in Fig. 5. highest formation redshift of the Phoenix series (z h = 1.17) is very well fit by both the Einasto and NFW profiles, with average residuals of only 3% and 6%, respectively. Indeed, a well defined correlation may be seen in Table 2 between quantitative measures of departures from equilibrium, such as the centre offset, d off, or the mass fraction in the form of substructure, f sub, and the average residuals from the best NFW and Einasto fits. On average, both indicators are substantially smaller for Aquarius than for Phoenix (Table 3), as expected. The higher formation redshift of galaxysized haloes means that they are closer to dynamical equilibrium than recently-assembled cluster haloes. 3.2 Pseudo-Phase-Space Density and Velocity Anisotropy The similarity in the mass profiles of galaxy- and cluster-sized CDM haloes highlighted in the previous subsection extends to their dynamical properties. We show this by comparing the spherically averaged pseudo-phase-space density (PPSD) profiles of Phoenix and Aquarius haloes. The PPSD, ρ/σ 3, is dimensionally identical to the phase-space density, but not strictly a measure of it. (Here the velocity dispersion, σ(r), is defined as the square root of twice the specific kinetic energy in each spherical shell.) It is well known that PPSD profiles are well approximated by a simple power-law, ρ/σ 3 r χ (Taylor & Navarro 2001), intriguingly similar to the secondary-infall self-similar solutions of Bertschinger (1985), where the exponent, χ = 1.875 (see also Ludlow et al. 2010, and references therein). PPSD profiles for all level-2 Phoenix clusters are shown in Fig. 7, and compared with the average PPSD for Aquarius haloes. Since clusters are denser and have higher velocity dispersions than galaxy-sized haloes, we scale all profiles to the scale radius, r 2, of each halo. Together with the density at this characteristic radius, ρ 2, these quantities define a characteristic velocity, V 2 = (Gρ 2 ) 1/2 r 2, that allows us to compare PPSD profiles of haloes of widely different mass in a meaningful way. The top panel of Fig. 7 shows that, in these scaled units, Aquarius and Phoenix have very similar PPSD profiles. The similarity extends over the range 0.06<r/r 2 < 4 where both simulation sets give converged results. (Note that Phoenix profiles actually probe radii interior to 0.06r 2 because of their lower concentration.) Table 3 lists the average parameters (and their dispersion) of power-law fits of the form ( ) ρ σ 3 = C ρ 2 r χ χ V 2 3, (3) r 2 where C χ = (σ(r 2 )/V 2 ) 3. Fits are carried out over the length range 0.06<r/r 2 < 4 for each halo. On average, both the slope and the normalisation of Aquarius and Phoenix PPSD profiles are almost indistinguishable emphasizing again the structural similarity between cluster- and galaxy-sized haloes. At the same time, the scatter is larger in the Phoenix sample than in Aquarius (Table 3), highlighting again the larger halo-tohalo variation of cluster profiles. This may also be appreciated in the bottom panel of Fig. 7, where residuals from the self-similar r 1.875 power law are shown. Although PPSD profiles scatter above and below the self-similar solution depending on the individual dynamical state of each cluster, the PPSD profile of cluster haloes seem to be, on average, indistinguishable from that of galaxy-sized haloes. We reach a similar conclusion when comparing the velocity anisotropy profiles of Phoenix clusters with those of Aquarius

10 Gao et al. 0.4 Ph-B-2 Ph-F-2 Ph-G-2 < Aquarius > 0.4 0.2 Ph-C-2 Ph-D-2 Ph-E-2 Ph-H-2 Ph-I-2 < Phoenix > 0.2 δ lnρ 0.0 δ lnρ 0.0-0.2-0.2 Einasto fits NFW fits -0.4 0.001 0.010 0.100 1.000 r/r 200-0.4 0.001 0.010 0.100 1.000 r/r 200 Figure 6. Residuals from the best Einasto (left panel) and NFW (right panel) profile fits for all level-2 Phoenix haloes. Colors and line types are as in Fig.5. The thick black dashed curve corresponds to the composite profile obtained after stacking all 9 Phoenix level-2 runs. The red thick dashed curve corresponds to the same composite profile, but for the 6 galaxy-sized level-2 Aquarius haloes. ) (ρ/σ 3 3 )/(ρ -2 /V -2 δln(ρ/σ 3 ) 10 5 10 4 10 3 10 2 10 1 10 0 10-1 10-2 0.4 0.2 0.0-0.2-0.4 Ph-B-2 Ph-C-2 Ph-D-2 Ph-E-2 Ph-F-2 Ph-G-2 Ph-H-2 Ph-I-2 <Phoenix> <Aquarius> 0.01 0.10 1.00 10.00 r/r -2 /2σ r 2 β=1-σ t 2 1.0 0.8 0.6 0.4 0.2 0.0-0.2 Ph-B-2 Ph-C-2 Ph-D-2 Ph-E-2 Ph-F-2 Ph-G-2 Ph-H-2 Ph-I-2 <Phoenix> <Aquarius> 0.01 0.10 1.00 10.00 r/r -2 Figure 7. Spherically averaged pseudo-phase-space density (PPSD; ρ/σ 3 ) of all level-2 Phoenix haloes as a function of radius. Profiles are plotted down to the convergence radius, r conv. Radii are given in units of the scale radius, r 2, of the best-fit Einasto profile for each halo. Densities are scaled to ρ 2 = ρ(r 2 ), and velocity dispersions, σ(r), to the characteristic velocity V 2 =(Gρ 2 ) 1/2 r 2. The thick dashed lines shows the average PPSD of all Phoenix (black) and Aquarius (red) haloes plotted over the converged radial range common to both simulation series: 0.06 (r/r 2 ) 4, respectively. The bottom panel shows residuals from a simple r 1.875 power-law fit. haloes (Fig. 8). Aside from a slightly larger scatter, the velocity anisotropy, which measures the ratio of the kinetic energy in tangential and radial motions, increases gently from the centre, where haloes are nearly isotropic, to the outer regions, where radial motions dominate. Phoenix and Aquarius haloes again seem indistinguishable from each other regarding velocity anisotropy when compared over their converged radial range. 3.3 Projected Profiles The preceding discussion highlights the mass invariance of the structure of CDM haloes, but it also makes clear that the dynamical Figure 8. Velocity anisotropy profiles of all level-2 Phoenix haloes. Radii are expressed in units of the scale radius, r 2, of the best-fit Einasto profile. Profiles are plotted down to the convergence radius, r conv. The thick dashed lines show the average anisotropy profile of all Phoenix (black) and Aquarius (red) haloes over the radial range where both give converged results (0.06 (r/r 2 ) 4). youth of clusters limits the validity of simple fitting formulae to describe their instantaneous mass profiles. This complication must be taken into account when comparing observational estimates of individual cluster mass profiles with the profiles expected in a CDMdominated Universe. Stacking clusters in order to obtain an average cluster profile might offer a way of circumventing this difficulty. This should smooth out local inhomogeneities in the mass distribution and average over different dynamical states to produce a more robust measure of the shape of the mass profile. Aside from dynamical youth, another issue that complicates the interpretation of observations is the fact that, due to the cluster s asphericity, projected mass profiles, such as those measured through gravitational lensing, may differ substantially from simple projections of the 3D spherically-averaged profiles discussed above. Depending on the line of sight a cluster may appear more or

The Phoenix project 11 less massive within a given radius, leading to biases in the cluster s estimated mass, concentration, and even the shape of its density profile. We illustrate this in Fig. 9, where we plot the surface density profile of two Phoenix clusters, and Ph-I-2, each projected along 20 different random lines of sight. The aspherical nature of the clusters results in large variations (up to a factor of 3) in the surface density in the inner regions. For comparison, we also show in Fig. 9 the result of a weak and strong-lensing analysis of a stack of four massive clusters by Umetsu et al. (2011). The mass of the stacked cluster lies between that of Ph-A and Ph-I, which explains why, on average, Ph-A Σ(R) profiles lie below the observed data whereas the opposite applies to Ph-I. In agreement with earlier work (see, e.g., Corless & King 2007; Sereno et al. 2010; Oguri et al. 2010, and references therein), Fig. 9 suggests that substantial biases may be introduced by projection effects on estimates of cluster parameters, especially when reliable data are restricted to the inner regions of a cluster. For example, fitting the inner 500h 1 kpc of the projected profile with an NFW profile results in mass and concentration (M 200, c) estimates that vary from (5.4 10 14 h 1 M, 4.8) to (7.3 10 14 h 1 M, 9.8) when using the projections that maximize or minimize the inner surface density, respectively (see Fig. 9). The corresponding numbers for Ph-I-2 are (1.8 10 15 h 1 M, 4.1) and (3.0 10 15 h 1 M, 7.1). Comparing these numbers with those listed in Tables 1 and 2 we see that variations as large as 30% in the mass and 60% in the concentration may be introduced just by projection effects 3. We explore this further in Fig. 10, where the small dots show the mass-concentration estimates for 500 random projections of each level-2 Phoenix cluster. Large symbols correspond to the 3D estimates listed in Tables 1 and 2. The black diamond symbol indicates the M 200 -c estimate for the stack of 4 strong-lensing clusters presented by Umetsu et al. (2011). This figure again emphasizes the importance of projection effects; for example, 12% of random projections result in concentration overestimates larger than 25%. Although an exhaustive analysis of such biases is beyond the scope of the present paper, the results in Figs. 9 and 10 suggest that there is no substantial difficulty matching the surface density profile of lensing clusters such as those studied by Umetsu et al. (2011). Our interpretation thus agrees with that reached by a number of recent studies (see, e.g., Oguri et al. 2011; Okabe et al. 2010; Gralla et al. 2011; Umetsu et al. 2011), which conclude that there is no obvious conflict between the concentration of lensing-selected clusters and those of ΛCDM haloes once projection effects are taken into account. Interestingly, despite the large variations in surface density alluded to above, the shape of the surface density profile is quite insensitive to projection effects. We show this in the righthand panels of Fig. 9; the weak dependence of γ p (R) on projection may thus be profitably used to assess the consistency of theoretical predictions with cluster mass profiles. For illustration, we compare in the same panels the logarithmic slope of the projected profile, γ p = d lnσ(r)/d lnr, with the stacked cluster data of Umetsu et al. (2011). Despite the fact that the mass of the simulated and observed clusters are different and that no scaling has been applied, there is clearly quite good agreement between observation and Phoenix clusters, supporting our earlier conclusion. Available data on individual clusters are bound to improve dramatically with the advent 3 Note that variations may actually be larger, because these estimates neglect the possible contribution of the large-scale mass distribution along the line-of-sight. log10(c) 1.0 0.8 0.6 0.4 0.2 Ph-E-2 Ph-B-2 Ph-F-2 Ph-C-2 Ph-G-2 Ph-I-2 Umetsu et al. (2011) Ph-D-2 Ph-H-2 14.6 14.8 15.0 15.2 15.4 log10(m 200 h/m O ) Figure 10. Cluster virial mass vs concentration estimated from fits to the projected density profiles of level-2 Phoenix haloes in the radial range R<500h 1 kpc. A total of 500 random projections are used for each halo. The large filled circles indicate the true value of the virial mass and concentration of the cluster, obtained from NFW fits to the 3D spherically-averaged profile (see Appendix and Table 2). The dashed curve flanked by dotted lines shows the fit to the mass-concentration relation derived by Neto et al. (2007). Note that projection effects lead to significant bias in the mass and concentration., which are underestimated on average by 8.5 ± 17% and 0.4±20%, respectively, where the error is the rms of all projections for the 9 clusters. The black diamond symbol indicates the M 200 -c estimate for a stack of 4 strong-lensing clusters taken from Umetsu et al. (2011). of surveys such as CLASH with the Advanced Camera for Surveys onboard the Hubble Space Telescope (Postman et al. 2011). These surveys will enable better constraints on the shape of the inner mass profile of individual rich clusters and it is therefore important to constrain how projection effects may affect them. Fig. 11 shows the distribution of γ p at two projected radii, R=3 and 10h 1 kpc. The histograms are computed after choosing 500 random lines of sight for each of our 9 level-2 Phoenix haloes. On average, cluster projected profiles flatten steadily towards the centre, from γ p =0.35 to 0.25 in that radial range, but with fairly large dispersion; the rms is σ γp = 0.054 and 0.091 at R=10 and 3h 1 kpc respectively. Because of the large dispersion it is unlikely that observations of a single cluster can lead to conclusive statements about the viability of CDM; however, it should be possible to use this constraint fruitfully once data for a statistically significant number of clusters become available. 4 THE SUBSTRUCTURE OF PHOENIX CLUSTERS As may be seen from the images presented in Fig. 3, substructure is ubiquitous in Phoenix clusters. We have used SUBFIND (Springel et al. 2001b) to identify and characterize self-bound structures (subhaloes) within the virial radius of the main halo. We discuss below the mass function, spatial distribution, and internal properties of subhaloes in Phoenix. Since our main goal is to explore the mass invariance of the properties of CDM haloes, we contrast these results with those obtained for the galaxy-sized Aquarius haloes.

12 Gao et al. 10 16 3.0 2.5 Σ [h M O Mpc -2 ] 10 15 10 14 Umetsu et al. 2011 γ p = -dlnσ/dlnr 2.0 1.5 1.0 NFW fits 0.5 10 13 0.01 0.10 1.00 R [h -1 Mpc] 0.0 0.01 0.10 1.00 R [h -1 Mpc] 3.0 10 16 Ph-I-2 2.5 Σ [h M O Mpc -2 ] 10 15 10 14 γ p = -dlnσ/dlnr 2.0 1.5 1.0 0.5 10 13 0.01 0.10 1.00 R [h -1 Mpc] 0.0 0.01 0.10 1.00 R [h -1 Mpc] Figure 9. Projected density profiles of (top) and Ph-I-2 (bottom). We show 20 different random projections for each cluster. The asphericity of the clusters leads to large variations (up to a factor of 3) in the projected density at a given radius depending on the line of sight. On the other hand, the shape of the profile (as measured by the logarithmic slope, γ p = d lnσ/d ln R, is much less sensitive to projection effects. Data with error bars correspond to the stacked profile of 4 massive clusters estimated using strong and weak lensing data (Umetsu et al. 2011). 10 6 10 6 10 5-1 N M sub 10 5-3.4 N V max 10 4 10 4 N M sub N 10 3 10 2 10 1 10 0 10 13 Ph-A-4 Ph-A-3 Ph-A-1 N 3 N V max 10 3 10 2 10 1 10 0 10 8 10 7 10 12 10 7 10 8 10 9 10 10 10 11 10 12 10 13 M sub [h -1 M O ] 10 6 10 100 V max [km s -1 ] Figure 12. Left: The cumulative mass function of substructure haloes ( subhaloes ) within the virial radius of cluster Ph-A at z = 0. We compare the results of four different realizations of the same halo, Ph-A-1 to Ph-A-4, with varying numerical resolution. The top and bottom panels contain the same information; the bottom shows the number of subhaloes weighted by mass or, equivalently, the fractional contribution of each logarithmic mass bin to the total mass in subhaloes. Each curve extends down to a mass corresponding to 60 particles. Note that, over the range resolved by the simulations, the cumulative function is well approximated by a power-law, N M 1, the critical dependence for logarithmically divergent substructure mass. Right: Same as left panels, but for the subhalo peak circular velocity.

The Phoenix project 13 0.3 10-1 0.2 3 kpc/h 10 kpc/h 10-2 dp/dγ p 0.1 r max [h -1 Mpc] 10-3 10-4 Subhalos (Ph-A-1) Isolated halos (Ph-A-1) Subhalos (Aq-A-1) Isolated halos (Aq-A-1) Neto et al. 2007 0.0 0.0 0.2 0.4 0.6 γ p = -dlnσ/dlnr 10-5 1 10 100 V max [km s -1 ] Figure 11. Distribution of the slope of the circularly-averaged surface density profile, γ p (R), measured at two different radii, R = 3 and 10h 1 kpc in projection. These histograms are based on 500 random lines of sight for each of the level-2 Phoenix clusters. Vertical arrows show the values corresponding to the projected profile of all nine clusters stacked together. The profiles become gradually shallower towards the centre, but with large scatter: γ p changes from 0.35 to 0.25 as R goes from 10 to 3h 1 kpc, but the halo-to-halo scatter is quite large, with rms of order 0.09 at 3h 1 kpc and 0.05 at 10h 1 kpc, respectively. Figure 14. Peak circular velocity, V max, vs the radius at which it is reached, r max. The solid cyan curve indicates the r max -V max relation obtained for isolated haloes in the Millennium Simulation by Neto et al. (2007). Subhaloes in both Phoenix (solid black curve) and Aquarius (solid red curve) deviate systematically from this relation towards smaller r max at a given velocity. This is a result of tidal stripping, which shifts the location of the peak inwards while changing little the peak velocity. Isolated haloes identified in Aquarius and Phoenix (shown with dashed lines) are not subject to tides and are in good agreement with the Millennium Simulation results. 4.1 Mass Function We start by analyzing the Ph-A simulation series in order to identify the limitations introduced by finite numerical resolution. The top left panel of Fig. 12 shows the cumulative mass function of subhaloes, N(> M), plotted in each case down to the mass corresponding to 60 particles. The bottom left panel shows the same data, but after weighting the numbers by subhalo mass, M sub, in order to emphasize the differences between runs. The results show clearly how, as resolution improves, the mass function converges at the low-mass end. Ph-A-4 agrees with higher resolution runs for subhaloes with mass exceeding 2 10 10 h 1 M, corresponding to roughly 150 particles; the same applies to Ph-A-3 for mass greater than 3 10 9 h 1 M, or 170 particles, and to for 7 10 8 h 1 M, or 140 particles. We conclude that the subhalo mass function can be robustly determined in Phoenix haloes down to subhaloes containing roughly 150 particles, in good agreement with the results reported for Aquarius haloes (see Fig. 6 of Springel et al. 2008a). For level-2 runs this implies a subhalo mass function that spans over 6 decades in mass below the virial mass of the halo. The subhalo mass function is also routinely expressed in terms of the subhalo peak circular velocity. This is shown in the right-hand panels of Fig. 12 which shows that level-2 Phoenix runs give robust estimates of the abundance of subhaloes down to V max 20 km s 1, a factor of 75 lower than the main halo s V 200. Both the subhalo mass and velocity functions seem reasonably well approximated by simple power laws: N Msub 1 and N V 3.4 max, respectively. Interestingly, the M 1 dependence corresponds to the critical case where each logarithmic mass bin contributes equally to the total mass in substructure. This is logarithmically divergent as M sub approaches zero, and implies that a significant fraction of the mass could in principle be locked in haloes too small to be resolved by our simulations. We note, however, that even at the resolution of Ph-A-1, of nearly 7 decades in mass, only 8% of the mass within r 200 is in the form of substructure. Extrapolating down to the Earth mass by assuming that N Msub 1, the total mass locked in substructure would still be only about 27 percent. Fig. 13 compares these results with other level-2 Phoenix clusters in order to assess the general applicability of the Ph-A subhalo mass function. The cumulative number of subhaloes N(> M) is weighted here by µ= M sub /M 200 (left panel) in order to emphasize differences as well as to enable the comparison of haloes of different virial mass. Although the subhalo mass function, expressed in this form, is relatively flat in several Phoenix clusters (indicative of an N Msub 1 dependence) it is clearly declining in others. The average trend, as indicated by the stacked Phoenix cluster (thick dashed black curve) may be approximated, in the range 10 6 < µ< 10 4, by N µ 0.98. This is a slightly steeper dependence than found for Aquarius haloes over the same mass range, N µ 0.94 (thick dashed red curve), but still subcritical. The slight difference in the average slope of the Aquarius and Phoenix subhalo mass functions is smaller than the halo-to-halo scatter in either simulation set. This is shown in Table 3, where we list the average parameters of power-law fits of the form, N(> µ)=n m (µ/10 6 ) s, (4) for Aquarius and Phoenix haloes. The dispersion around s is similar to the difference between the average slope of Aquarius and Phoenix haloes, suggesting that there is no significant difference in the shape of the subhalo mass function of cluster- and Milky-way halo-sized haloes. Fig. 13 also shows that substructure is slightly more prevalent in clusters than in galaxy-sized haloes. Indeed, at all values of M sub /M 200 the number of Phoenix subhaloes exceeds that in Aquarius, and this is reflected in the higher values of N m (7866 for Phoenix vs 5092 for Aquarius; see Table 3). This is another consequence of the dynamical youth of clusters compared to galaxies (tides take a few orbital times to strip a subhalo), as may be verified by inspection of Table 1: in the cluster that forms latest, Ph-G, sub-